You are on page 1of 8

Relative dating is the science of determining the relative order of past events (i.e.

,
the age of an object in comparison to another), without necessarily determining
their absolute age, (i.e. estimated age). In geology, rock or superficial deposits,
fossils and lithologies can be used to correlate one stratigraphic column with
another. Prior to the discovery of radiometric dating which provided a means of
absolute dating in the early 20th century, archaeologists and geologists used this
technique to determine ages of materials. Though relative dating can only
determine the sequential order in which a series of events occurred, not when they
occur, it remains a useful technique especially in radiometric dating. Relative dating
by biostratigraphy is the preferred method in paleontology, and is in some respects
more accurate (Stanley, 16769). The Law of Superposition, which states that older
layers will be deeper in a site than more recent layers, was the summary outcome
of 'relative dating' as observed in geology from the 17th century to the early 20th
century.
The regular order of occurrence of fossils in rock layers was discovered around 1800
by William Smith. While digging the Somerset Coal Canal in southwest England, he
found that fossils were always in the same order in the rock layers. As he continued
his job as a surveyor, he found the same patterns across England. He also found
that certain animals were in only certain layers and that they were in the same
layers all across England. Due to that discovery, Smith was able to recognize the
order that the rocks were formed. Sixteen years after his discovery, he published a
geological map of England showing the rocks of different geologic time eras.
Principles of relative chronology
Methods for relative dating were developed when geology first emerged as a formal
science. Geologists still use the following principles today as a means to provide
information about geologic history and the timing of geologic events.
Uniformitarianism
The principle of Uniformitarianism states that the geologic processes observed in
operation that modify the Earth's crust at present have worked in much the same
way over geologic time.[1] A fundamental principle of geology advanced by the
18th century Scottish physician and geologist James Hutton, is that "the present is
the key to the past." In Hutton's words: "the past history of our globe must be
explained by what can be seen to be happening now."[2]
Intrusive relationships
The principle of intrusive relationships concerns crosscutting intrusions. In geology,
when an igneous intrusion cuts across a formation of sedimentary rock, it can be
determined that the igneous intrusion is younger than the sedimentary rock. There
are a number of different types of intrusions, including stocks, laccoliths, batholiths,
sills and dikes.
Cross-cutting relationships
The principle of cross-cutting relationships pertains to the formation of faults and
the age of the sequences through which they cut. Faults are younger than the rocks
they cut; accordingly, if a fault is found that penetrates some formations but not
those on top of it, then the formations that were cut are older than the fault, and
the ones that are not cut must be younger than the fault. Finding the key bed in
these situations may help determine whether the fault is a normal fault or a thrust
fault.[3]
Inclusions and components
The principle of inclusions and components states that, with sedimentary rocks, if
inclusions (or clasts) are found in a formation, then the inclusions must be older
than the formation that contains them. For example, in sedimentary rocks, it is
common for gravel from an older formation to be ripped up and included in a newer
layer. A similar situation with igneous rocks occurs when xenoliths are found. These
foreign bodies are picked up as magma or lava flows, and are incorporated, later to

cool in the matrix. As a result, xenoliths are older than the rock which contains
them.
Original horizontality
The principle of original horizontality states that the deposition of sediments occurs
as essentially horizontal beds. Observation of modern marine and non-marine
sediments in a wide variety of environments supports this generalization (although
cross-bedding is inclined, the overall orientation of cross-bedded units is horizontal).
Superposition
The law of superposition states that a sedimentary rock layer in a tectonically
undisturbed sequence is younger than the one beneath it and older than the one
above it. This is because it is not possible for a younger layer to slip beneath a layer
previously deposited. The only disturbance that the layers experience is
bioturbation, in which animals and/or plants move things in the layers. however,
this process is not enough to allow the layers to change their positions. This
principle allows sedimentary layers to be viewed as a form of vertical time line, a
partial or complete record of the time elapsed from deposition of the lowest layer to
deposition of the highest bed.
Faunal succession
The principle of faunal succession is based on the appearance of fossils in
sedimentary rocks. As organisms exist at the same time period throughout the
world, their presence or (sometimes) absence may be used to provide a relative age
of the formations in which they are found. Based on principles laid out by William
Smith almost a hundred years before the publication of Charles Darwin's theory of
evolution, the principles of succession were developed independently of
evolutionary thought. The principle becomes quite complex, however, given the
uncertainties of fossilization, the localization of fossil types due to lateral changes in
habitat (facies change in sedimentary strata), and that not all fossils may be found
globally at the same time.[4]
Lateral continuity
The principle of lateral continuity states that layers of sediment initially extend
laterally in all directions; in other words, they are laterally continuous. As a result,
rocks that are otherwise similar, but are now separated by a valley or other
erosional feature, can be assumed to be originally continuous.
Layers of sediment do not extend indefinitely; rather, the limits can be recognized
and are controlled by the amount and type of sediment available and the size and
shape of the sedimentary basin. Sediment will continue to be transported to an area
and it will eventually be deposited. However, the layer of that material will become
thinner as the amount of material lessens away from the source.
Often, coarser-grained material can no longer be transported to an area because
the transporting medium has insufficient energy to carry it to that location. In its
place, the particles that settle from the transporting medium will be finer-grained,
and there will be a lateral transition from coarser- to finer-grained material. The
lateral variation in sediment within a stratum is known as sedimentary facies.
If sufficient sedimentary material is available, it will be deposited up to the limits of
the sedimentary basin. Often, the sedimentary basin is within rocks that are very
different from the sediments that are being deposited, in which the lateral limits of
the sedimentary layer will be marked by an abrupt change in rock type.
Inclusions of igneous rocks
Melt inclusions are small parcels or "blobs" of molten rock that are trapped within
crystals that grow in the magmas that form igneous rocks. In many respects they
are analogous to fluid inclusions. Melt inclusions are generally small - most are less
than 100 micrometres across (a micrometre is one thousandth of a millimeter, or
about 0.00004 inches). Nevertheless, they can provide an abundance of useful
information. Using microscopic observations and a range of chemical microanalysis

techniques geochemists and igneous petrologists can obtain a range of useful


information from melt inclusions. Two of the most common uses of melt inclusions
are to study the compositions of magmas present early in the history of specific
magma systems. This is because inclusions can act like "fossils" - trapping and
preserving these early melts before they are modified by later igneous processes. In
addition, because they are trapped at high pressures many melt inclusions also
provide important information about the contents of volatile elements (such as H2O,
CO2, S and Cl) that drive explosive volcanic eruptions.
Sorby (1858) was the first to document microscopic melt inclusions in crystals. The
study of melt inclusions has been driven more recently by the development of
sophisticated chemical analysis techniques. Scientists from the former Soviet Union
lead the study of melt inclusions in the decades after World War II (Sobolev and
Kostyuk, 1975), and developed methods for heating melt inclusions under a
microscope, so changes could be directly observed.
Although they are small, melt inclusions may contain a number of different
constituents, including glass (which represents magma that has been quenched by
rapid cooling), small crystals and a separate vapour-rich bubble. They occur in most
of the crystals found in igneous rocks and are common in the minerals quartz,
feldspar, olivine and pyroxene. The formation of melt inclusions appears to be a
normal part of the crystallization of minerals within magmas, and they can be found
in both volcanic and plutonic rocks.
Included fragments
The law of included fragments is a method of relative dating in geology. Essentially,
this law states that clasts in a rock are older than the rock itself.[5] One example of
this is a xenolith, which is a fragment of country rock that fell into passing magma
as a result of stoping. Another example is a derived fossil, which is a fossil that has
been eroded from an older bed and redeposited into a younger one.
This is a restatement of Charles Lyell's original principle of inclusions and
components from his 1830 to 1833 multi-volume Principles of Geology, which states
that, with sedimentary rocks, if inclusions (or clasts) are found in a formation, then
the inclusions must be older than the formation that contains them. For example, in
sedimentary rocks, it is common for gravel from an older formation to be ripped up
and included in a newer layer. A similar situation with igneous rocks occurs when
xenoliths are found. These foreign bodies are picked up as magma or lava flows,
and are incorporated, later to cool in the matrix. As a result, xenoliths are older than
the rock which contains them.
Archaeology
Relative dating methods in archaeology are similar to some of those applied in
geology. The principles of typology can be compared to the biostratigraphic
approach in geology.
Linguistics
Relative dating is often employed in historical linguistics, most typically in study of
historical phonology and of loanwords.
Planetology
Relative dating is used to determine the order of events on objects other than Earth;
for decades, planetary scientists have used it to decipher the development of
bodies in the Solar System, particularly in the vast majority of cases for which we
have no surface samples. Many of the same principles are applied. For example, if a
valley is formed inside an impact crater, the valley must be younger than the crater.
Craters themselves are highly useful in relative dating; as a general rule, the
younger a planetary surface is, the fewer craters it has. If long-term cratering rates
are known to enough precision, crude absolute dates can be applied based on

craters alone; however, cratering rates outside the Earth-Moon system are poorly
known.(Hartmann, 258)

Multiple melt inclusions in an olivine


crystal. Individual inclusions are oval
or round in shape and consist of
clear glass, together with a small
round vapor bubble and in some
cases a small square spinel crystal.
The black arrow points to one good
example, but there are several
others. The occurrence of multiple

inclusions within a single crystal is


relatively common

You might also like